Sie sind auf Seite 1von 18

atmosphere

Review
Assessment of Indoor-Outdoor Particulate Matter Air
Pollution: A Review
Matteo Bo 1 ID
, Pietro Salizzoni 2 , Marina Clerico 1 and Riccardo Buccolieri 3, * ID

1 Politecnico di Torino, DIATI—Dipartimento di Ingegneria dell’Ambiente, del Territorio e delle Infrastrutture,


corso Duca degli Abruzzi 24, Torino 10129, Italy; matteo.bo@polito.it (M.B.); marina.clerico@polito.it (M.C.)
2 Laboratoire de Mécanique des Fluides et d’Acoustique, UMR CNRS 5509 University of Lyon,
Ecole Centrale de Lyon, INSA Lyon, Université Claude Bernard Lyon I, 36, avenue Guy de Collongue,
Ecully 69134, France; pietro.salizzoni@ec-lyon.fr
3 Dipartimento di Scienze e Tecnologie Biologiche ed Ambientali, University of Salento,
S.P. 6 Lecce-Monteroni, Lecce 73100, Italy
* Correspondence: riccardo.buccolieri@unisalento.it; Tel.: +39-0832-297-062

Received: 9 June 2017; Accepted: 20 July 2017; Published: 26 July 2017

Abstract: Background: Air pollution is a major global environmental risk factor. Since people
spend most of their time indoors, the sole measure of outdoor concentrations is not sufficient to
assess total exposure to air pollution. Therefore, the arising interest by the international community
to indoor-outdoor relationships has led to the development of various techniques for the study
of emission and exchange parameters among ambient and non-ambient pollutants. However,
a standardised method is still lacking due to the complex release and dispersion of pollutants
and the site conditions among studies. Methods: This review attempts to fill this gap to some extent
by focusing on the analysis of the variety of site-specific approaches for the assessment of particulate
matter in work and life environments. Results: First, the main analogies and differences between
indoor and outdoor particles emerging from several studies are briefly described. Commonly-used
indicators, sampling methods, and other approaches are compared. Second, recommendations for
further studies based on recent results in order to improve the assessment and management of those
issues are provided. Conclusions: This review is a step towards a comprehensive understanding of
indoor and outdoor exposures which may stimulate the development of innovative tools for further
epidemiological and multidisciplinary research.

Keywords: indoor-outdoor; mass concentration; nanoparticles; particle number concentration (PNC);


PM10 ; PM2.5 ; sampling; Total Suspended Particles (TSP); ultrafine particles (UFP)

1. Introduction
In many countries, the persistence or the increasing of air pollution represents a major
environmental and health issue [1], which largely depends on the amount of chemical energy used in
our society (i.e., fossil, biomass). The relapses of the anthropic activities cannot be related exclusively
to local emissions in urban and metropolitan areas, but also to the diffuse pollution involving entire
territories or mega-city regions [2]. This is the case, for example, of the Po Valley in Italy, where the
urban emission contribution is overlapped with the critical state of pollution at the regional scale, in
particular during wintertime periods.
The assessment of source emissions and the measuring and modelling of outdoor concentrations is,
therefore, fundamental to obtain a framework of pollution conditions of an area at different temporal
and spatial scales. International and national legislation and policies are mainly based on these
approaches. Despite a general improvement of observation and measurement techniques of outdoor
pollutant concentrations in the last 20 years, due to technological developments and the adoption of

Atmosphere 2017, 8, 136; doi:10.3390/atmos8080136 www.mdpi.com/journal/atmosphere


Atmosphere 2017, 8, 136 2 of 18

some normative restrictions, the implemented policies and actions have shown limitations in reducing
personal exposure [3]. While the policy-makers at various public entities’ scales are challenged for the
introduction of innovative actions, the scientific community is called to make a step forward in the
assessment of air pollution and its relapses to different targets and in different environments.
The correlation among outdoor air pollution and health diseases, affecting in particular respiratory
and cardiovascular systems, has been widely demonstrated [2,4–6]. However, an approach based
exclusively on the assessment of outdoor air pollution has shown its limited effectiveness. People
spend, in fact, most of their time indoors [7,8] and the correlation among personal exposure and
outdoor concentrations of particulate matter is still weak in the literature [9,10]. For this reason, despite
formerly and recent epidemiological studies referring mainly to outdoor particulate concentrations [11],
the assessment of indoor and personal concentrations in work and life environments is necessary to
evaluate the total exposure to air pollution.
Furthermore, direct health diseases are primary in a wide list of relapses of air pollution which
also include disturbances to the population and the loss in quality and in the use of territories and
indoor environments. This is even more serious especially in highly-populated areas, where the
synergy of air pollution with other hazard factors (i.e., noise, vibrations, odours) may lead to increasing
damages and disturbance to human health and land use [12,13]. Even though in most epidemiological
studies the assessment of finer particulate sizes prevails, the employ of other indicators to understand
the whole phenomena affecting human health and the use of environments is required.
Within this context, the present review analyses the existing works on the assessment of
indoor-outdoor (I-O) particulate matter concentrations and relationships. Differently from other
existing reviews on I-O particulate matter pollution [14–16], this work is not limited to investigate
one specific parameter or approach, but studies which considered more indicators (i.e., Total
Suspended Particles (TSP), PM10 , PM2.5 , PM10–2.5 , ultrafine particles (UFP), Nanoparticles (NP), and
Indoor/Outdoor Ratio (I/O ratio), air exchange rate, infiltration factors) in residential and working
environments are preferred to find potentials and weaknesses in the framework of I-O PM research
methodologies. Such recent studies, in fact, proposed sampling or modelling approaches for the
assessment of site-specific cases. The lack of standardised methods, due to the complex phenomenology
of air pollution release and dispersion and the boundary conditions, is evident. This review attempts
to partially fill such a gap.

2. Materials and Methods


Starting from the definition of the main objectives, this review collects a large dataset of papers
based on several main searches of key criteria which include: large-scale international studies
on indoor-outdoor air pollution issues, on-site assessment of I-O concentrations in specific life
environments and work sectors, and on-site and experimental studies of particulate matter with
different size between indoor and outdoor environments.
The dataset is then reduced by specific exclusion criteria. The main focus is toward studies
that analysed more than one indicator; however, studies focusing on one specific indicator are also
considered. Likewise, studies involving different indoor and outdoor environments, such as residences
and workplaces, are considered. Furthermore, recent studies are generally preferred and, in particular,
those proposing innovative approaches and new points of views on methodologies and results; studies
published prior to 2005 are also considered for their significant contribution to following studies or
that focused on “atypical” case studies. Large-scale exposure assessment studies developed at the end
of the last century, such as EXPOLIS, PTEAM, and THEES [9,17–19], are intended as starting points for
the purpose of this review.
Using the SciVal tool [20], a qualitative analysis for investigating the main tendencies in I-O
studies is also performed. For the searching criteria, indicator terms associated to indoor and outdoor
keywords was compiled in order to observe their usage in recent years (2011–2016 published papers).
The analysis was developed both by considering the overall results from the SciVal DB and then by
Atmosphere 2017, 8, 136 3 of 18

filtering to journal categories (JC) expressing subject areas (SA), such as: “Environmental Science”;
“Earth and Planetary Sciences”; “Engineering”; and “Medicine” (each journal could be characterised
by more than one subject area). The research was developed in May 2017.

3. Brief Summary of the Main Characteristics of PM in Indoor and Outdoor Environments


The main sources of outdoor PM pollution in most developed countries are commonly identified to
be road traffic (including exhaust and non-exhaust emissions of vehicle combustion, tire wearing, and
resuspension), power generation plants, industries, agriculture, and domestic heating systems [2,21].
While natural sources, which represent a consistent fraction of aerosols in many regions, contribute
mainly on coarser particles, anthropic sources are well-known for the generation of primary and
secondary fine, ultrafine, and nano-scale particulates [22–25]. The definition and adoption by
normative frameworks of size-depending indicators to fix mass concentration limit values for outdoor
air quality considered both the penetration in the human respiratory tract and the need to distinguish
anthropogenic and natural emissions [26].
Indoor sources are associated to anthropic activities and the intended use of spaces. In life
environments a significant role is played by smoking and cooking, followed by heating systems,
cleaning, and resuspension due to the presence of humans [10,27–30]. Combustion processes and
cleaning contribute significantly to fine, ultrafine, and nanoparticles emissions, while coarse fractions
of PM are principally evidenced due to resuspension [22,24,27,31,32]. In working environments, PM
size distribution, concentrations, and chemical properties are even more site-specific than in residential
ones as these depend on the used materials, productive methods, and working typologies. Extensive
literature on school environments, partially for assessing children’s exposure (i.e., the tendency for
health impairments, and the large percentage of daytime spent in those spaces) and as a major working
sector for the number of employees, is found [14,33–35].
A large amount of works described the consistent contribution of outdoor PM to indoor
concentrations. The heterogeneity in the estimation of such contributions found for different particulate
size ranges is strictly linked to the different pathways of infiltration and aerodynamic behaviours of
finer and coarse particles. A general trend describing a decreased penetration for coarser particles is
found in the literature [27]. Other factors influencing the contribution of outdoor pollution on indoor
environments are constituted by the type of ambient ventilation (i.e., natural or mechanical), distance to
the sources, meteorological conditions, and by the building age and architectural characteristics [28,31].
Indoor versus outdoor levels are found to be heterogenic in the literature. In the absence of intense
indoor sources, studies show a general trend of higher outdoor concentrations rather than indoor
values [8,36,37]. Furthermore, spatial and temporal variability of outdoor PM could significantly affect
the relation between I-O concentrations [38]. In urban areas, as an example, the source proximity and
the primary and secondary particulate pathways lead to the high variability of the observed PM size
distributions and concentrations. An even more uniform spatial distribution is generally observed for
the finer particles, and frequent exceptions are reported in the literature due to the interference of local
sources [26,39]. Moreover, the seasonal variability of I-O relationships may be referred mainly to the
outdoor contribution, the influence of ventilation types, and occupant behaviours [26].
A widely-used indicator for evaluating the indoor-outdoor exchange is the ratio between the
measured concentrations in those environments [40,41]. The results of I/O ratios varied in the literature
from values tending to zero in modern mechanically-ventilated buildings with an absence of indoor
activities to values over 10 in the occupational sector or for residential buildings with intense indoor
activity and smoking in territories with relatively low outdoor concentrations [10,37,42]. This wide
range is consistently influenced by resuspension, air exchange, and the deposition velocity of particles.
Therefore, this indicator does not permit reaching a complete explanation of I-O relationships.
In the literature [15,40,43,44] various sets of parameters and models which consider the
mechanisms of generation (i.e., the contribution of indoor sources), transport (i.e., the air exchange
rate, infiltration factor, penetration factor, change in indoor concentration per unit change in outdoor
Atmosphere 2017, 8, 136 4 of 18

concentration) ,and deposition (i.e., the decay rate) of particulate are presented. Evidence from the
literature frameworks reinforce the review hypothesis of the need for methods which consider different
indicators and parameters, rather than limiting the assessment to the ones required by the normative
framework or for the assessment of specific patterns.

4. Indoor-Outdoor Particulate Matter Sampling and Assessment

4.1. Indicators
Within the reviewed papers, the analysis of the main indicators used to describe particulate
matter mass and particle number concentrations (PNC) is developed. Both standard and non-standard
indicators found in I-O studies are considered.

4.1.1. Total Suspended Particles (TSP)


Total Suspended Particles (TSP) is a historically-used indicator for the assessment of outdoor
air pollution and its relapses. However, a limited employ emerges in the literature for the purposes
of I-O assessment. This outcome follows the general trend of substitution of such indicators, due
to the availability of consolidated technologies and techniques, with others taking into account the
size distribution of airborne particulate (i.e., PM10 , PM2.5 ) in outdoor applications. Moreover, even
if the interest on assessing TSP for intense natural and human dust emissions persists, such an
indicator has a limited effectiveness for observing outdoor particle infiltration phenomena into indoor
ambient environments.
The few recent uses of TSP are mainly related to the determination of specific components of
particulates. Nazir et al. [45] assessed I-O distributions of trace metal in the TSP of outdoor origins
(i.e., industry, vehicles, soils sources). As expected, higher values of TSP are found outdoors, rather
than indoors, with moderate correlation among the two environments (R = 0.415). Similar results are
found in other studies [46,47] which investigate I-O concentrations of particle phase PAHs in total dust.
Other applications are found in researches considering large sets of indicators including TSP, PM10 ,
PM4 , PM2.5 , PM1 , and UFP at different environments, such as museums, offices, industries, schools,
and an Antarctic research station [48–52].

4.1.2. PM10 and PM2.5


In the last two decades there has been an extensive and increasing use of PM10 and PM2.5 . These
indicators are adopted in a variety of different I-O studies and are considered as main objects of the
samplings or as references for the comparison within other indicators or with other studies.
Ranges of PM10 and PM2.5 mass concentrations vary by orders of magnitude between, and
within, indoor and outdoor environments. In the month of June in two modern offices with mechanical
ventilation, Chatoutsidou et al. [53] report low indoor daily averaged PM10 concentrations (<3.5 µg/m3 )
while simultaneously-collected outdoor measures ranged between 11 µg/m3 and 21 µg/m3 . Higher
outdoor than indoor concentrations are also reported by Diapouli et al. [54] at three residences with
air-conditioning in the Athens urban area. However, they found approximately ten times higher
24 h-averaged PM10 indoor values (≈25–47 µg/m3 ) compared to the results of Chatoutsidou et al. [53].
Additionally, in two naturally-ventilated commercial activities, higher mean PM10 indoor values
(≈50–55 µg/m3 ) rather than outdoor (≈25–45 µg/m3 ) are found by Vicente et al. [55] over the sampling
period. Same authors also found higher indoor concentrations of PM10 during working hours rather
than non-working hours. A similar temporal variation between occupancy and non-occupancy is also
found in a previous study by Branis et al. [56] which, however, developed the study using different
sampling strategies and time references as discussed later in this review. High variability of PM2.5 I-O
concentrations is also described by other analysed studies [40,50,57–61]. In particular, the findings of
Liu et al. [58] through different residential and commercial buildings in Beijing, clearly show the wide
variability within indoor PM10 and PM2.5 concentrations which are, resultantly, higher in restaurants,
Atmosphere
Atmosphere 2017,
2017, 8,
8, 136
136 55 of
of 18
18

commercial buildings in Beijing, clearly show the wide variability within indoor PM10 and PM2.5
dormitories, and
concentrations classrooms,
which rather than
are, resultantly, in supermarkets,
higher in restaurants, computer
dormitories,rooms, andoffices, and libraries
classrooms, rather(PMthan 10
and PM2.5 ranging, respectively, from 373.8 µg/m 3 and 136.6 µg/m3 in restaurants to 33.8 µg/m3 and
in supermarkets, computer rooms, offices, and libraries (PM10 and PM2.5 ranging, respectively, from
373.8 μg/m3 3in
5.6 µg/m andlibraries).
136.6 μg/m3 in restaurants to 33.8 μg/m3 and 5.6 μg/m3 in libraries).
In the literature, less frequently-adopted PM classes
In classes of of indicators
indicators (PM (PM55, PMPM44, PM22,, PM11)) are are
likewise investigated. The use of such indicators is, in many cases,
likewise investigated. The use of such indicators is, in many cases, related to the cut-off of the related to the cut-off of the
availability of
availability of instrumentation,
instrumentation, the the purposes
purposes of of the
the occupational
occupational field field investigations,
investigations, or or for
for evidence
evidence
of specific
of specific size
size fraction
fraction emissions.
emissions. In In aa former
former study,
study, Monn Monn et et al.
al. [8] investigated
investigated I-O I-O and
and personal
personal
relationships by
relationships by comparing
comparing indoor and outdoor outdoor PM PM10 10 and PM2.5 measurements with
2.5 measurements with PM55 personal
personal
samplings. The
samplings. Theuse use of
of PM
PM44isisfound
foundin inWeichenthal
Weichenthal et et al.
al. [62]
[62] where
where the measure of indoor PM44 in in
relation I-O measurements
relation to I-O measurementsofofultrafine
ultrafineparticles
particles(UFP)
(UFP) is is reported.
reported. MoreMore recently,
recently, Diapouli
Diapouli et al.et[54]
al.
[54] monitored
monitored I-O I-O
massmassandand
PNPN concentrations
concentrations of of UFP,
UFP, black
black smoke,
smoke, PM PM , ,and
1010 andPMPM22,, the
the latter using
using
aa custom-made
custom-made impactorimpactor (with
(with a cut-off point at at 2.12.1 µm μm and and at at 23
23 L/min).
L/min). The The useuse of
of PM11 is is
documented in three studies at schools and universities of Central Europe and studies assessing
documented in three studies at schools and universities of Central Europe and in studies assessing
multiple
multiple indicators
indicators (including
(including TSP, PM10 10,, and
and PM
PM2.52.5) )[51,56–58,63,64].
[51,56–58,63,64].
A
A summary
summary of of the
the observed
observed 24-h24-h mean
mean concentrations
concentrations in in indoor
indoor and and outdoor
outdoor environments
environments is is
reported
reported in Figure 1, which shows the box box plot
plot for
for PM
PM10 10 and PM2.5 2.5..The
Theselection
selectionof ofsuch
such indicators
indicators is is
ascribed
ascribed toto the
the availability
availability of extensive data in the reviewed studies.

Figure 1. Box
Figure 1. Boxplot
plot ofh-mean
of 24 24 h-mean
values values
of PM10of
andPM10 and
PM PM32.5
2.5 [µg/m ] in[μg/m
indoor3] and
in outdoor
indoor environments.
and outdoor
environments.
4.1.3. Ultrafine Particles (UFP) and Nanoparticles (NP)
4.1.3. Ultrafine Particles (UFP) and Nanoparticles (NP)
Despite the limited amount of paper reporting the use of UFP compared to PM10 and PM2.5 ,
Despiteinterest
the recent the limited
in theseamount
typesof of
paper reporting
particles in I-O thestudies
use of UFP
is wellcompared
documented to PMin 10 and PM2.5, the
the literature.
recent
The main content of the studies on UFP particles is related to the investigation of PNC andThe
interest in these types of particles in I-O studies is well documented in the literature. main
relapses
content of the studies on UFP particles is related to the investigation of PNC
to human health. The dominant contribution of such particles to number concentrations is, in fact, and relapses to human
health. The dominant
contraposed contribution
to the incidence of such
of larger size particles
particles to to number concentrations
mass concentrations [26].is, in fact, contraposed
to theWeichental
incidence ofetlarger
al. [62]sizeinvestigated
particles to massthe concentrations
contribution to[26]. UFP from heating systems in life
Weichental et al. [62] investigated the
environments (in over 36 houses) taking into account outdoor contribution to concentrations,
UFP from heating systems
the age in life
of buildings,
environments
types of cooking (insystems,
over 36 houses) taking into
and the presence account outdoor
of smokers. Results ofconcentrations,
the study describe the agetheof buildings,
contribution
types of cooking systems, and the presence of smokers. Results of the
of these sources in UFP indoor concentration, in particular in the residential buildings using electric study describe the
contribution of these sources in UFP indoor concentration, in particular in
baseboard heaters and wood stoves. However, as also reported by the authors, the ambient conditionsthe residential buildings
using
and theelectric baseboard
potential dominant heaters and wood
contribution stoves.
of other However,
sources as also
(outdoor, reported
cooking, by the authors,
and smoking) need tothe
be
ambient conditions and the potential dominant contribution of other sources
taken into account in the comparison of the measured UFP concentrations in buildings with different (outdoor, cooking, and
smoking) need toIncreasing
heating systems. be taken into account
interest in theinuse
theofcomparison of the ismeasured
the UFP indicator reported UFP concentrations
for specific in
workplaces,
buildings with different heating systems. Increasing interest in the use of the UFP indicator is
Atmosphere 2017, 8, 136 6 of 18

such as offices and copy centres, where significant indoor sources of fine and ultrafine particles are
represented by laser printers [55,58,65,66].
Concentrations vary from low values (<103 particles/cm3 ) in particularly clean indoor
environments [37] to high values (>106 particles/cm3 ) in the presence of intense indoor sources
of such particles [67]. Diapouli et al. [68] report higher UFP mean concentrations outdoors
(32,000 ± 14,200 particles/cm3 ) rather than indoors (24,000 ± 17,900 particles/cm3 ), with a maximum
indoor mean value found in a library with a carpet-covered floor and a smoking office
(both ≈52,000 particles/cm3 ). Similar results are also found in other studies [53,60,61]. In particular,
Zauli Sajani et al. [61] investigate I-O concentrations in the front and back of a building along a high traffic
street: the highest UFP 1 h-mean value is found for the outdoor front sampling (≈25,500 particles/cm3 )
and the lowest in the indoor located in the back of the building (≈3500 particles/cm3 ). Comparable
values are found indoors at the front and outdoors at the back (7635 particles/cm3 and 7444 particles/cm3 ,
respectively) leading the authors to suggest similarities between the gradients of front-back (I-O)
and of high-low traffic areas (I-O). Seasonal variability investigated by Wheeler et al. [60] show
both indoor and outdoor 24 h-averaged values are higher during winter rather than summer.
Furthermore, the difference between indoor and outdoor concentrations is narrow during one of
the sampled periods (the second summer campaign) due to a significantly lower outdoor averaged
value compared to an another considered period with the same awaited meteorological conditions
(first summer campaign).
Related to nanoparticles, relatively few studies focused directly on the assessment of I-O
relationships. However, a growing interest is found in occupational indoor environments due to
the increase of productive activities employing innovative nanomaterials [69]. Dahm et al. [70], for
example, investigate the exposure to carbon nanotubes and nanofibers in six productive sites (handling
materials with diameter ranges between 1.1 nm and 140 nm). In this research, no evident trends are
described for mass and PN concentrations among the different factories and within the I-O samplings
by using three different real-time optical instruments and a filter-based method.

4.1.4. Miscellaneous
In the reviewed articles, a heterogeneity in the use of multiple indicators is observed: the joint
use of two indicators is found in different studies with PM10 and PM2.5 or PM2.5 and UFP, while
studies considering more than two indicators employed PM10 and PM2.5 with UFP or TSP [38,60,61].
The recurrence of the use non-typical PM classes (i.e., PM5 , PM4 , PM1 ) is also found in many studies,
frequently joint with the most-used indicators [49,50,54,57,62,71].
One of the main direct results of the studies assessing more indicators is the investigation of
correlations and contributions among different particle size ranges. Mass concentration of finer sizes
appear to contribute variously to the values of greater size classes. Liu et al. [58] showed a significant
correlation and contribution of PM10 to TSP (R2 = 0.674–0.996, range 47–69%) and discrete for PM2.5 to
PM10 (R2 = 0.144–0.894, range 16–45%) and PM1 to PM10 (R2 = 0.149–0.879, range 6–23%). In another
study, Branis et al. [56] show a higher correlation of PM2.5 and PM10 by comparing indoor PM2.5 /PM10
ratio during workdays (R = 0.872 daytime; R = 0.912 nighttime) and weekends (R = 0.918 daytime;
R = 0.991 nighttime). High correlation between indoor PM1 /PM10 and PM1 /PM2.5 with a slighter
increase during the weekend is also observed. The presence of a source of coarse particles during
workdays, which does not appear during the other observed periods and can be associated to a good
correlation among PM10–2.5 and student presence (R = 0.683), can be attributable to the resuspension.
Similar results are provided by Vicente et al. [55], who show ratios between different size classes
ranging from 0.62 to 0.78 for PM2.5 /PM10 and PM4 /PM10 during working hours and tending to
unity during the hours of non-occupancy, as an effect of coarse particle decay. The same study
reports slight unity values for PM1 /PM2.5 and PM10 /TSP, both for working and non-working periods.
A correlation among indoor concentrations of PM4 and UFP (R2 = 0.53) is reported in the I-O research
of Weichental et al. [62].
Atmosphere 2017, 8, 136 7 of 18
Atmosphere 2017, 8, 136 7 of 18
4.2. Sampling Methods
Based onMethods
4.2. Sampling the available systematic reviews on particulate air sampling technologies [72,73], it is
worth examining instrumentation and approaches commonly employed in the reviewed I-O
Based on the available systematic reviews on particulate air sampling technologies [72,73], it is worth
investigations.
examining
Related instrumentation
to concentration and samplings,
approaches gravimetric
commonly employed
methods in arethethe
reviewed
most used I-O investigations.
(Figure 2). The
reasonRelated
is due to to
concentration
the historical samplings,
development gravimetric
of thismethods are the
technology, themost used (Figure
normative 2). Theand
references, reason
the
is due to the historical development of this technology, the normative references,
need to collect mass on filters for further analysis (i.e., the composition of PM). In particular, and the need to collect
mass on filters
gravimetric for further
cyclones andanalysis
impactors, (i.e.,such
the composition of PM). In particular,
as the widely-employed Harvard gravimetric
Impactor cyclones
[74], areand
the
impactors, such as the widely-employed
most used technologies for collecting mass. Harvard Impactor [74],
The microbalances are
(i.e., the most
TEOM, QCM)used technologies
appear for
marginally
collecting
consideredmass. in I-OThestudies,
microbalances (i.e., TEOM,
with applications QCM) appear
restricted in timemarginally
sampling,considered in I-O studies,
assessed environments,
with
used indicators, or limited to the validation of other instrumentation [75–77]. Relating tolimited
applications restricted in time sampling, assessed environments, used indicators, or to the
the material
validation
of membranes,of other instrumentation
Teflon filters are the [75–77].
most used,Relating to the by
followed material
quartzofandmembranes, Teflon
glass filters. Thefilters are the
diameter of
most used, followed by quartz and glass filters. The diameter of membranes
membranes are typically represented by 25 mm, 37 mm, and 47 mm filters, with the smaller are typically represented
by 25 mm, 37associated
membranes mm, and 47with mm personal
filters, with the smaller
samplers membranes
as, for example, associated with personal samplers
personal environmental monitors
as,
(PEM) reported by Meng et al. [40]. Pumping flow rates in the studies range between 2.3 L/min andflow
for example, personal environmental monitors (PEM) reported by Meng et al. [40]. Pumping 38.3
rates
L/minin(with
the studies
a highrange between
frequency of 102.3L/min
L/min andand 38.3
16.7 L/min
L/min) (with
due a high frequency
to instrument designofand 10 L/min and
compliance
16.7
withL/min) due[78].
legislation to instrument design and compliance with legislation [78].

Figure 2. The employed instrument technologies for the assessment of mass and particle number
Figure 2. The employed instrument technologies for the assessment of mass and particle number
concentrations in the reviewed studies.
concentrations in the reviewed studies.

An alternative approach to obtain approximate mass concentrations is represented by the use of


An alternative approach to obtain approximate mass concentrations is represented by the use
real-time techniques, such as optically-based systems. The availability of “instantaneous” results
of real-time techniques, such as optically-based systems. The availability of “instantaneous” results
and higher time resolution than gravimetric methods represents an important key factor for the
and higher time resolution than gravimetric methods represents an important key factor for the
sampling campaign design. Advantages comprise the possibility to develop short-term and spot
sampling campaign design. Advantages comprise the possibility to develop short-term and spot
approaches at different locations and extract trends and behaviours over long-term monitoring. The
approaches at different locations and extract trends and behaviours over long-term monitoring.
use of such instrumentations should be subjected to dedicated site-specific verifications in order to
The use of such instrumentations should be subjected to dedicated site-specific verifications in
investigate the need of correction factors to reduce bias with concentrations obtained using
order to investigate the need of correction factors to reduce bias with concentrations obtained using
gravimetric methods [73,79]. Additionally, the main use of real-time instruments in I-O studies is
gravimetric methods [73,79]. Additionally, the main use of real-time instruments in I-O studies is
related to the collection of PNC and size distribution. Chatoutsidou et al. [53] developed
related to the collection of PNC and size distribution. Chatoutsidou et al. [53] developed measurements
measurements of mass (PM10) and number concentrations in recently-built offices using real-time
of mass (PM10 ) and number concentrations in recently-built offices using real-time instruments:
instruments: I/O simultaneously measures are collected by light scattering photometric
I/O simultaneously measures are collected by light scattering photometric instrumentation (PM10 ),
instrumentation (PM10), scanning mobility particle sizer spectrometers (SMPS, for particles under 0.7
scanning mobility particle sizer spectrometers (SMPS, for particles under 0.7 µm), and aerodynamic
μm), and aerodynamic particle sizer spectrometers (APS, for particles amid 0.5–18 μm). Results
particle sizer spectrometers (APS, for particles amid 0.5–18 µm). Results show both mass and PN
show both mass and PN concentrations to be higher outdoors rather than indoors (I/O < 0.3 for all
concentrations to be higher outdoors rather than indoors (I/O < 0.3 for all the measures), with PNC
the measures), with PNC dominated by finer particles (the concentrations of particles with a
Atmosphere 2017, 8, 136 8 of 18

diameter lower than 0.5 μm over two orders of magnitude higher than the range 0.5–18 μm) and
dominated
mass by finer
by coarser particles
particles (the concentrations
(related, in particular, to of the
particles
human with a diameter
presence lower than
in working hours0.5 µm over
leading to
resuspension). Despite air conditioning determined to cause a significant reduction of PMin
two orders of magnitude higher than the range 0.5–18 µm) and mass by coarser particles (related, 10
particular, to thetemporal
concentrations, human presence
fluctuationsin working
indoor hours
are foundleading to resuspension).
comparable to outdoor Despite
PM10 air conditioning
variability over
determined to cause a significant reduction of PM10 concentrations, temporal fluctuations indoor are
time.
found comparable
Most to outdoor PMare
of the measurements 10 variability
made by over fixedtime.
instruments. However, personal samplers are
used to investigate the personal exposure with on-board However,
Most of the measurements are made by fixed instruments. personal samplers
direct measurements. are studies
These used to
describe the need to follow dedicated design criteria to ensure the validity of such approachesneed
investigate the personal exposure with on-board direct measurements. These studies describe the to
as, for
follow dedicated
example, measuringdesignat criteria
fixedto ensure
positions the validity
in bedrooms of such approaches as, for example,
during nighttime measuringthe
and keeping at
fixed positions in bedrooms during nighttime and keeping the instrumentation
instrumentation far from high-humidity sources [80]. Moreover, such sampling methods should be far from high-humidity
sources
easy to [80].
carryMoreover,
and should such sampling
integrate methods
noise control should
and be easyinsulation
sound to carry and shouldinintegrate
systems order tonoisenot
control and
interfere with sound
personalinsulation
daytime systems
activitiesin order to not interfere
and, therefore, with personal daytime
the representativeness activitiesdata.
of the collected and,
therefore, the representativeness
Technological developments improved of the collected data. Technological
such issues compared todevelopments
heavy personal improved
samplers such issues
used in
compared to heavy personal samplers used in former studies as Koistinen
former studies as Koistinen et al. [81]. The use of personal samplers for developing measurements at et al. [81]. The use of personal
samplers
fixed for developing
position measurements
is also reported. The goodatcorrelation
fixed position is also
of this reported.
sampling The good
solution withcorrelation
traditionaloffixed
this
sampling solution with traditional fixed monitoring samplers found
monitoring samplers found by Meng [40] suggests a sufficient accuracy of such instruments for by Meng [40] suggests a sufficient
accuracy ofI/O
describing such instrumentsThe
relationships. for describing
feasibility of I/O therelationships.
methodologies Therequires,
feasibility of the methodologies
however, a site-specific
requires, however, a site-specific testing
testing activity in parallel with traditional approaches.activity in parallel with traditional approaches.
The characteristic
The characteristic time time interval
intervalof ofsampling
samplingisisfound foundtotobebepredominantly
predominantly ofof2424h or 4848
h or h (Figure
h (Figure3).
TheThe
3). daily
dailyinterval
interval is is
widely
widely adopted
adoptedtotofollow followand andcompare
comparethe theresults
results with
with outdoor
outdoor normative
normative
prescriptions which are frequently based on daily concentrations
prescriptions which are frequently based on daily concentrations from midnight to midnight. from midnight to midnight. Custom
24-h samplings
Custom with starting
24-h samplings with and ending
starting andpoints
ending at different
points at hoursdifferentof the dayof
hours arethereported
day aredue to the
reported
campaign design requirements. The sampling interval of 48 h is
due to the campaign design requirements. The sampling interval of 48 h is typically related to typically related to epidemiological
studies. Similar percentages
epidemiological studies. Similar can percentages
be found on can the review
be found ononPM the by Mohammed
2.5 review on PM2.5 et byal. [14]. Among
Mohammed et
the[14].
al. otherAmong
intervals, thethe 8 h intervals,
other sampling periodthe 8 hissampling
associated with the
period typical working
is associated with the time and, therefore,
typical working
is adopted
time and, on I-O studies
therefore, developed
is adopted on atI-O
workplaces rather thanat
studies developed in workplaces
life environments. rather Furthermore,
than in life
environments. Furthermore, spot measurements (<8 h for each sampling position)general
spot measurements (<8 h for each sampling position) are reported to investigate trends to
are reported of
pollutants or
investigate due totrends
general limitedofavailable
pollutants time to spend
or due for in-site
to limited sampling.
available time toMoreover,
spend forain-site
formersampling.
study by
Branis et al. [56] adopted a custom 12 h time interval in order to collect
Moreover, a former study by Branis et al. [56] adopted a custom 12 h time interval in order to collect mass concentrations on filters
describing
mass different periods
concentrations on filtersof use of a university
describing different hall (lecture
periods ofhours
use ofvs. nighttime).hall
a university Among long-term
(lecture hours
measurement campaigns, semi-continuous approaches are
vs. nighttime). Among long-term measurement campaigns, semi-continuous approaches are alsoalso reported [60]. Such methodologies
do not permit
reported the collection
[60]. Such methodologiesof representative
do not permit datathe over the timeof
collection series, but merelydata
representative to have
overathegeneral
time
picturebut
series, of the phenomena
merely to have aover the sampling
general picture ofperiod.the phenomena over the sampling period.

Figure 3. Sampling intervals adopted in the reviewed studies.


Figure 3. Sampling intervals adopted in the reviewed studies.
Atmosphere 2017, 8, 136 9 of 18

Most of the reviewed studies are developed using simultaneous sampling intervals between
indoorAtmosphere 2017, 8, 136
and outdoor measurements. The design of contemporary measurement campaigns is9 of 18
adopted
in order to neglect the time variability of PM among I-O environments. Few exceptions are reported in
Most of the reviewed studies are developed using simultaneous sampling intervals between
relation to limited
indoor parts ofmeasurements.
and outdoor the studies (i.e.,Theadditional
design ofsamplings,
contemporary instrumentation
measurement failures)
campaigns or for
is the
unfeasibility of achieving simultaneous samplings as, for example, due to
adopted in order to neglect the time variability of PM among I-O environments. Few exceptions are extreme weather conditions
reported in theinAntarctica
reported relation tostudy limitedbyparts
Pagelofetthe
al. [49].
studies (i.e., additional samplings, instrumentation
In accordance
failures) or for thewith good-practice
unfeasibility and normative
of achieving simultaneous frameworks,
samplings as, all for
theexample,
studies dueset the instrument
to extreme
weather conditions reported in the Antarctica study by Pagel et al. [49].
chains far from every potentially interfering source. For this reason, indoor sampling is typically made
In accordance
at the centre with good-practice
of the environment and at aand normative
height between frameworks,
1 and 2 m all(to
thesimulate
studies set the
the instrument
breathing height
chains far from every potentially interfering source. For this reason,
of occupants). Distances from walls and heating sources are also identified as main design indoor sampling is typically
factors.
made at the centre of the environment and at a height between 1 and 2 m (to simulate the breathing
Related to outdoor air samplings, in studies the height ranges from 1–2 m (referred to the front door or
height of occupants). Distances from walls and heating sources are also identified as main design
ground level) to the height of the floor corresponding to the indoor sampling (5–10 m from the ground).
factors. Related to outdoor air samplings, in studies the height ranges from 1–2 m (referred to the
Measures
front on
doorbuilding
or ground roofs aretoalso
level) the reported [57,79].
height of the floor The horizontaltodistance
corresponding the indoor from the external
sampling (5–10 m wall is
frequently
from the reported
ground).as anotheron
Measures key parameter
building for also
roofs are the reported
reduction of unwanted
[57,79]. interference.
The horizontal Personal
distance from
samplers are typically positioned in the breathing zone of the carrying
the external wall is frequently reported as another key parameter for the reduction of unwantedperson.
In some studies
interference. mainly
Personal related
samplers aretotypically
epidemiological
positionedpurposes [36,50,56]
in the breathing zonethe outdoor
of the concentrations
carrying person.
are totally,Inor some
partially,studies
obtainedmainly
fromrelated to epidemiological
central-site stations, instead purposes
of proximity [36,50,56]
samplings. the The
outdoor
adoption
concentrationsfrom
of concentrations are the
totally,
publicor partially,
monitoring obtained
stations from central-site
or other samplings stations, instead an
represents of opportunity
proximity to
samplings. The adoption of concentrations from the public monitoring stations
save resources. However, this requires both the detailed examination of representativeness of the station or other samplings
represents an opportunity to save resources. However, this requires both the detailed examination of
for describing outdoor values at the indoor sampling and the comparison of measurement methods.
representativeness of the station for describing outdoor values at the indoor sampling and the
When possible, it is strongly recommended to perform measurements of outdoor concentrations in
comparison of measurement methods. When possible, it is strongly recommended to perform
proximity to the indoor
measurements measure.
of outdoor The choice in
concentrations of proximity
the samplers’ disposition
to the indoor measure.is strictly
Therelated
choice to ofthe
thegoals
of thesamplers’
survey and requires an accurate aprioristic design, as underlined by Zauli
disposition is strictly related to the goals of the survey and requires an accurate aprioristic Sajani et al. [61].
design, as underlined by Zauli Sajani et al. [61].
4.3. Studies of Site-Specific and Environmental Characteristics
4.3. Studies of Site-Specific and Environmental Characteristics
As expected, different sites and ambient characteristics emerge through the reviewed literature.
First, the As expected,
type different
of analysed sites and ambient
environments: characteristicsnumber
a comparable emerge through the reviewed
of case-studies literature.
is found between
First, the type of analysed environments: a comparable number of case-studies
workplaces (productive, non-productive) and life environments (Figure 4). Residential homes appearis found between
as theworkplaces (productive,
most represented, non-productive)
followed by schools andand
lifeuniversities,
environmentsoffices,
(Figureand4). commercial
Residential homes
buildings.
appear as the most represented, followed by schools and universities, offices, and commercial
Only a few I-O studies considered industries [50,70]. The choice to investigate different typologies of
buildings. Only a few I-O studies considered industries [50,70]. The choice to investigate different
buildings is also reported [50,58] and clearly indicates the interest of the international community to
typologies of buildings is also reported [50,58] and clearly indicates the interest of the international
develop I-O studies
community on a wide
to develop range on
I-O studies of alife andrange
wide occupational
of life and environments.
occupational environments.

Figure 4. Synthesis of the environment typologies assessed in the reviewed studies.


Figure 4. Synthesis of the environment typologies assessed in the reviewed studies.
Atmosphere 2017, 8, 136 10 of 18

Related to the spatial variability, urban and suburban areas prevail over rural areas. Most studied
indoor life environments are kitchens and living rooms due to the long time spent by occupants
and to the presence of the major residential indoor sources (cooking, smoking). In the occupational
sector, offices and ambient conditions potentially subjected to intense indoor sources (i.e., printers
in commercial copy centres) are investigated. The design of indoor measurements in unoccupied
residential environments is also found due to epidemiological purposes [61]. Outdoor environments
are frequently chosen as the front door and front windows of indoor ambient environments, followed
by courtyards and roofs.
Another difference among the studies is represented by the seasonal and duration variability
of assessment campaigns. The choice of seasonal periods is based principally on outdoor sampling
design both for meteorological constraints (i.e., rainfall, temperature) and for the need of measuring
during high- or low-pollution periods. Many studies analyse more different periods of the year
covering cold and hot seasons. While a predominance of a specific month or season do not emerge,
frequently multi-seasonal durations are represented by autumn-winter-spring campaigns for the
assessment of I-O relationships during such a particularly-sensible period for outdoor PM pollution.
Spring-autumn and winter-summer investigations are also frequently adopted as representative of
antipode meteorological conditions. These observations are based on the campaign duration reported
by the authors of the original works: the effective sampling intervals could involve only a part of
the entire period of study. As a brief observation, samplings are generally conducted over the entire
campaign period, or at least on multiple weeks/months, rather than limited to single days or weeks.
No prevalence of mechanically- or naturally-ventilated buildings emerges from the studies.
The type of ventilation assessed by the studies is, in fact, mainly related to the available sampling
areas, with the exception to studies on the effects of a specific ventilation design to the infiltration
of outdoor particles in indoor environments [53,79,82]. Significant differences are shown, in fact, in
relation to ventilation systems, as described in Section 5.

5. Exchange Factors

5.1. Indoor/Outdoor Ratio


For the purposes of this review, the term “exchange factors” denotes the indicators describing
the I-O relationship’s characteristics. The I/O ratio is widely used, even though it presents some
limitations, as described in Section 3. In particular, its high variability also emerges among the reviewed
studies and no consistent global trends can be stated a priori.
A general relationship with indoor activities and I/O ratio higher than the unity is found by
Diapouli et al. [68] which report PM10 and PM2.5 ratios higher than 1 in gym, offices, and classes, and
lower than 1 for a library (used only for a limited part of the day by students). Similar results are
obtained for PM10 by Vicente et al. [55] by observing I/O ratio at copy centres in workdays (mostly > 1,
with a maximum of 2.38) and in weekends (≈0.7–0.8). Significantly high I/O ratios for TSP, PM10 ,
PM2.5 , and PM1 (ranging from values approx. between 2 and 18, except for PM1 between 0.98 and 8.9)
are found at an Antarctic research station [49]. In this case study, the extreme environmental conditions
and limited outdoor sources significantly affect the increase of the observed I/O ratios, which should
depend on the indoor activities and the emissions of the research station (i.e., vehicles, incinerator).
The maximum observed I/O ratio (30.40), reached in an air conditioned classroom during cleaning
hours and in the presence of rainfall, is reported by Guo et al. [82]. I/O ratios less than unity is found
in the published studies of either occupied or unoccupied indoor environments [55,57,58,61,64,68,82].
A significant example of the indicator variability can be clearly observed in the results of
Challoner et al. [79]. The two highest I/O values (9.18 and 8.18), compared to other values between
1 and 3, are found for two offices (A and B) with consistent differences among them and some
“contradictory” results from the expected trends: naturally (A) and mechanically (B) ventilated (no
significant difference found in this study by building ventilation types); at the fifth (A) and ground (B)
Atmosphere 2017, 8, 136 11 of 18

Atmosphere 2017, 8, 136 11 of 18

floors(B)
(expected different
floors (expected outdoor
different concentration
outdoor concentrationincidence);
incidence); small
smalloffice
office(A)
(A) and largeopen
and large openplan
plan (B)
(incidence of air volume and ventilation pathways to indoor contribution);
(B) (incidence of air volume and ventilation pathways to indoor contribution); both during both during non-working
hoursnon-working
(even if during
hoursworking
(even if hours
duringhigher
working values
hoursare generally
higher valuesexpected rather
are generally than non-working,
expected rather than in
this case they are inferior
non-working, and they
in this case correspond to values
are inferior of 4.68 andto2.87,
and correspond respectively);
values and respectively);
of 4.68 and 2.87, ventilation intake
andofventilation
in front intake
a high-traffic in front
road of (with
for (B) a high-traffic road for
the expected (B) (with
effect the expected
of consistent effectconcentrations
outdoor of consistent on
outdoor concentrations on reducing the ratio). The study also developed
reducing the ratio). The study also developed two different campaigns for (B) considering two different campaigns
a different
disposition of the outdoor sampling: an I/O ratio four times higher is obtained withhigher
for (B) considering a different disposition of the outdoor sampling: an I/O ratio four times is
the outdoor
obtained with the outdoor station at the ground floor (8.18) than on the roof (2.04), where outdoor
station at the ground floor (8.18) than on the roof (2.04), where outdoor concentrations are expected
concentrations are expected to be lower. Considering the need of detailed correlation analysis with
to be lower. Considering the need of detailed correlation analysis with the indoor sources and the
the indoor sources and the characteristic meteorological conditions of the area (i.e., temporal
characteristic
variation,meteorological
well-noted highconditions
humidity, of thefrequent
and area (i.e., temporal
rainfalls variation,
in the Dublin well-noted high
region), these humidity,
results
and frequent rainfalls
confirm the extremein susceptibility
the Dublin region), these
of the I/O results
ratio confirmconditions.
to boundary the extreme susceptibility
Such of the I/O
variability among
ratio the
to boundary
reviewed conditions. Such variability
studies is synthetized among
in Figure the reviewed
5, which shows the studies
box is synthetized
plot reporting in Figure 5,
mean,
which shows the
maximum, box
and plot reporting
minimum values ofmean, maximum, and minimum values of I/O ratios.
I/O ratios.

Figure
Figure 5. Variability
5. Variability of mean,
of mean, maximumand
maximum and minimum
minimum I/O
I/Oratios in the
ratios reviewed
in the papers.
reviewed papers.

5.2. Air Exchange Rate (AER)


5.2. Air Exchange Rate (AER)
Evaluation of building ventilation is needed for the comprehension of influence of outdoor air
Evaluation
pollution intoofindoor
building ventilation
environments. is needed
Using for the comprehension
the air exchange of influence
rate (AER), expressed as the airof outdoor
flow rate air
pollution
per volume of the indoor environment, a minor exchange of air is reported for rate
into indoor environments. Using the air exchange rate (AER), expressed as the air flow
per volume of the indoor environment,
mechanically-ventilated a minorthan
buildings rather exchange of air is reported
naturally-ventilated for mechanically-ventilated
[40,60,62] ones. Volume
buildings rather
estimation than naturally-ventilated
is obtained [40,60,62]
in the studies by direct ones.
and indirect Volumeasestimation
approaches measures ofis obtained
indoor spaces,in the
cadastre
studies data,and
by direct and indirect
reported approaches
data by occupants. Measures
as measures of of the AER
indoor are carried
spaces, using
cadastre tracer
data, and gasses
reported
or the pressurized blower method. Uses of perfluorocarbon, carbon dioxide, and
data by occupants. Measures of the AER are carried using tracer gasses or the pressurized blower perfluorinated
methyl-cyclohexane
method. are reported [40,60,61,82].
Uses of perfluorocarbon, The distance
carbon dioxide, of tracer sourcesmethyl-cyclohexane
and perfluorinated from ventilation or are
heating sources is needed in order to reduce their interference. Different results for mechanical and
reported [40,60,61,82]. The distance of tracer sources from ventilation or heating sources is needed in
natural ventilations are found due to the age of the building, occupants’ behaviours, and the
order to reduce their interference. Different results for mechanical and natural ventilations are found
efficiency of filters. The greater seal in mechanically-ventilated buildings is generally associated to a
due to the penetration
lower age of the building, occupants’
of PM of outdoor originbehaviours, and the efficiency
than for naturally-ventilated of filters.[54].
environments The greater seal
in mechanically-ventilated buildings is generally associated to a lower penetration of PM of outdoor
origin than for naturally-ventilated environments [54].
Atmosphere 2017, 8, 136 12 of 18

5.3. Other Approaches


Among the I-O reviewed studies, considerations on some exchange factors or modelling
approaches, rather than the sole analysis of the I/O ratio, are reported. Meng et al. [40] adopted
a single compartment mass balance model and a random component superposition statistical model
(RCS), with the support of AER and PM direct measurements and values obtained from the literature,
to evaluate the contribution of ambient and non-ambient sources to indoor and personal concentrations.
Results showed an average contribution of outdoor pollution to PM2.5 indoor concentrations over
60%. Furthermore, the increasing trend of the infiltration factor (IF) at the growth of AER values
reported in the research shows the general reduction of outdoor contributions due to the decrease
of air exchange. Using a two-compartment mass balance model, Chatoutsidou et al. [66] have also
estimate the incidence of printer emissions in different rooms in a mechanically-ventilated building.
Within the results of the study published by Hoek et al. [38] emerges the role of particle size
and composition to their infiltration and penetration in indoor environments: using a single mass
balance model and comparing the results with other studies, the incidence of both dimensions and
chemical components is confirmed by the authors from the variability of the average IF among
sulphate, soot, and mass and particulate number concentrations (whole range between 0.06 and 0.87).
The assessment of sulphate in parallel with the other I-O multiple samplings, is also found in other
reviewed studies [50,68,83]. Actually, indoor sulphate concentration is a good estimator of IF due to the
absence of significant indoor sources of such inorganic pollutants and constitutes, considering some
limitations, a surrogate for describing the contribution of outdoor sources to indoor environments, as
described in the literature [84].
As previously stated, the description in detail of infiltration and penetration factors, as well as
modelling approaches, is intentionally considered marginal for the purposes of this review. Indicators
and models for describing the dynamics of the unwanted entrance of outdoor pollution into indoor
environments are systematically evaluated in other reviews [15,22,31,84]. We suggest to the readers
who aim to obtain a complete framework of the subject to refer to such, and to other works in the
existing literature.

6. Discussion
The framework emerging from the review of the existing literature shows a growing interest on
PM I-O assessment. Most of recent studies focus on the evaluation of fine and ultrafine particulates
due to the relevance of the relapses of potentially highly-penetrating particles to human health.
Additionally, the use of a standard indicator, such as the PM10 , is widely documented and the adoption
of innovative or non-standard indicators is also observed in recent studies. The observed outline from
the analysis of indicators is representative of the variety of approaches adopted within the I-O studies.
With the aim of finding potential standard approaches for the design and execution of I-O
assessment, the need to associate a common indicator to the ones adopted for the specific objective of
the studies is strengthened. This may represent an opportunity to make comparisons among studies,
considering prior research. The feasibility of such additional measurement with a standard indicator must
be subjected to the projects’ available resources and technologies. For this reason, the adoption of PM10
or PM2.5 additional sampling and the description of the I-O ratio could represent a good compromise
between the use of resources and having a general comparison of indoor and outdoor environments
among studies. Furthermore, the use of other indicators can also generally lead to a better comprehension
of indirect relapses of PM as an annoyance and disturbance or to evaluate a specific contribution to mass
concentrations, as a resuspension of dust, by assessing coarse particles (i.e., PM10–2.5 ).
In order to design a measurement campaign, the site-specific conditions of indoor and outdoor
environments are not of secondary importance. The potential role of different sources should be
evaluated: for outdoor sources the distance within the receptors, the emissions over time, and
the dispersion pathways should be considered while, for indoor environments, human activities
and behaviours (i.e., smoking, cooking, presence of pets) and internal and external architectural
Atmosphere 2017, 8, 136 13 of 18

characteristics (i.e., age of the building, carpets, doors) represent crucial parameters for the definition
of the assessment procedures [15,36,68]. In addition, sampling disposition should take into account
the interference with unwanted sources of heat, ventilation, and other pollutants. Simultaneous
measurements and outdoor sampling in the proximity of indoor environments should also be included
as design criteria. The aim is to reduce spatial and temporal variability, which is widely described for
indoor and outdoor PM. Contemporaneity is found to be generally considered in the reviewed studies,
while a persistence of studies adopting outdoor values from other samplings is observed.
Related to human exposure, which is one of the main goals of the reviewed studies, the weak
potential of using the sole outdoor concentrations to represent personal exposure to PM is confirmed
by the literature. The good correlation found among indoor workplaces and personal samplings
during working hours or, likewise, indoor residential and leisure time personal concentrations [85],
show the feasibility of a simplified approach-based fixed monitoring in different I-O environments to
obtain personal exposures [60]. Furthermore, the choice to associate I-O measurement with personal
samplings represents a major opportunity: a well-designed personal measurement can describe
unintended individual behaviours and might cover specific exposures to PM which can be difficult
to assess by adopting only fixed monitoring. An example is represented by in-vehicle exposures
which can consistently contribute to the “personal cloud” [9] for people spending a large amount
of daytime moving from one indoor environment to another (i.e., the workplace and home) or
working in transportation sectors. The recent developments of low-cost, light, and real-time samplers,
used in association with GPS technologies, represent an opportunity for further studies employing
personal measurements as shown by Steinle et al. [77]. The effectiveness of personal concentration
measurements, however, requires as a critical parameter of design the verification of the aptitude of
participants to carry the sampler, as also reported by Meng et al. [40]. Definitely, the design of both
fixed and personal approaches is suggested.
The good precision and the correlation with traditional fixed monitoring stations, demonstrated
in the literature [40], suggest the feasibility of using personal samplers in fixed positions for some
specific purposes. In fact, the use of personal samplers might represent a good alternative to assess
indoor and outdoor concentrations instead of traditional approaches, as described by Snyder et al. [86].
Different scenarios comprehend limitations in instrumentation and resources, difficulties of carrying
heavy samplers and reaching the investigated site, the need of short-time or spot sampling, and the
unavailability of connection to the electricity grid. It is, however, important to underline that the good
correlation founded in some studies [40] does not necessarily imply a complete correspondence of
the obtained concentrations and requires a reliable choice of sampling parameters considering the
site-specific boundary conditions.
As a final consideration, the available literature showed the wide interest in I/O studies on the
assessment of particulate matter chemical components between indoor and outdoor environments, in
different contexts, especially schools [34,45,49,61,68,83]. In particular, a broad investigation of PM-bound
PAHs is widely reported by recently-published papers [87–90] and a dedicated review on such topics is
under consideration by the authors, given the increasing interest of the scientific community.

7. Conclusions
This paper reviews recent studies on the assessment of I-O particulate matter air pollution.
An increasing interest by the international scientific community to indoor-outdoor relationships is
evidenced by the development of several techniques for the study of emission and exchange parameters
among ambient and non-ambient pollutants. The present review emphasises the importance of
reducing divergences among I-O studies, derived from differences in measuring approaches, site
characteristics, and campaign periods, and by the definition and adoption of standard approaches. This
statement is joint with the need of efforts for the definition of dedicated normative frameworks, with
the implementation of international common policies and specific value limits for indoor environments.
Atmosphere 2017, 8, 136 14 of 18

The opportunity given by the recent technological developments and the need to assess the
human exposure to recently-introduced materials and in less studied environments are continuously
leading to a growing interest in the development of innovative epidemiological and multidisciplinary
studies in the field of I-O assessment.

Author Contributions: Matteo Bo, Pietro Salizzoni, Marina Clerico, and Riccardo Buccolieri conceived and
designed the structure of the paper. Matteo Bo conducted the literature research, analysed the data, and wrote the
main part of the paper. All authors contributed to the discussion of the results and have read and approved the
final manuscript.
Conflicts of Interest: The authors declare no conflict of interest.

Abbreviations
I-O indoor-outdoor
Mass concentration concentration expressed in micrograms per cube meters [µg/m3 ]
NP nanoparticles
PM (airborne) particulate matter
PN Particle number
Particle number concentration = concentration expressed in number of particles
PNC
per cubic centimetre [particles/cm3 ]
UFP ultrafine particles

References
1. World Health Organization (WHO). Ambient Air Pollution: A Global Assessment of Exposure and Burden
of Disease. Available online: http://www.who.int/phe/publications/air-pollution-global-assessment/en/
(accessed on 11 May 2017).
2. International Agency for Research on Cancer (IARC). International Agency for Research on Cancer (IARC)
Monographs on the Evaluation of Carcinogenic Risks to Humans, Volume 109; Outdoor Air Pollution; IARC:
Lyon, France, 2015; Available online: http://monographs.iarc.fr/ENG/Monographs/vol109/index.php
(accessed on 12 July 2016).
3. World Health Organization (WHO). WHO Guidelines for Indoor Air Quality: Selected Pollutants.
2010. Available online: http://www.euro.who.int/en/health-topics/environment-and-health/air-quality/
publications/2010/who-guidelines-for-indoor-air-quality-selected-pollutants (accessed on 12 May 2017).
4. Brunekreef, B.; Holgate, S.T. Air pollution and health. Lancet 2002, 360, 1233–1242. [CrossRef]
5. Raaschou-Nielsen, O.; Andersen, Z.J.; Beelen, R.; Samoli, E.; Stafoggia, M.; Weinmayr, G.; Hoffmann, B.;
Fischer, P.; Nieuwenhuijsen, M.; Brunekreef, B.; et al. Air pollution and lung cancer incidence in 17 European
cohorts: Prospective analyses from the European Study of Cohorts for Air Pollution Effects (ESCAPE).
Lancet Oncol. 2013, 14, 813–822. [CrossRef]
6. Seaton, A.; Godden, D.; MacNee, W.; Donaldson, K. Particulate air pollution and acute health effects. Lancet
1995, 345, 176–178. [CrossRef]
7. Brasche, S.; Bischof, W. Daily time spent indoors in German homes—Baseline data for the assessment of
indoor exposure of German occupants. Int. J. Hyg. Environ. Health 2005, 208, 247–253. [CrossRef] [PubMed]
8. Monn, C.; Fuchs, A.; Högger, D.; Junker, M.; Kogelschatz, D.; Roth, N.; Wanner, H.U. Particulate matter less
than 10 µm (PM10 ) and fine particles less than 2.5 µm (PM2.5 ): Relationships between indoor, outdoor and
personal concentrations. Sci. Total Environ. 1997, 208, 15–21. [CrossRef]
9. Ozkaynak, H.; Xue, J.; Spengler, J.; Wallace, L.; Pellizzari, E.; Jenkins, P. Personal exposure to airborne particles
and metals: Results from the Particle TEAM study in Riverside, California. J. Expo. Anal. Environ. Epidemiol.
1996, 6, 57–78. [PubMed]
10. Wallace, L. Indoor Particles: A Review. J. Air Waste Manag. Assoc. 1996, 46, 98–126. [CrossRef] [PubMed]
11. Chen, R.; Kan, H.; Chen, B.; Huang, W.; Bai, Z.; Song, G.; Pan, G. Association of Particulate Air Pollution with
Daily Mortality: The China Air Pollution and Health Effects Study. Am. J. Epidemiol. 2012, 175, 1173–1181.
[CrossRef] [PubMed]
Atmosphere 2017, 8, 136 15 of 18

12. Bo, M.; Clerico, M.; Pognant, F. Annoyance and disturbance hazard factors related to work and life
environments: A review. Geam-Geoing. Ambient. E Mineraria-Geam-Geoengin. Environ. Min. 2016, 149, 27–34.
13. Kephalopoulos, S.; Koistinen, K.; Paviotti, M.; Schwela, D.; Kotzias, D. Proceedings of the
International Workshop on “Combined Environmental Exposure: Noise, Air Pollutants and Chemicals”.
Available online: https://ec.europa.eu/jrc/en/publication/eur-scientific-and-technical-research-reports/
proceedings-international-workshop-combined-environmental-exposure-noise-air-pollutants-and (accessed on
10 May 2017).
14. Mohammed, M.O.A.; Song, W.-W.; Ma, W.-L.; Li, W.L.; Ambuchi, J.J.; Thabit, M.; Li, Y.-F. Trends in
indoor–outdoor PM2.5 research: A systematic review of studies conducted during the last decade (2003–2013).
Atmos. Pollut. Res. 2015, 6, 893–903. [CrossRef]
15. Chen, C.; Zhao, B. Review of relationship between indoor and outdoor particles: I/O ratio, infiltration factor
and penetration factor. Atmos. Environ. 2011, 45, 275–288. [CrossRef]
16. Lin, C.-C.; Peng, C.-K. Characterization of Indoor PM10 , PM2.5 , and Ultrafine Particles in Elementary School
Classrooms: A Review. Environ. Eng. Sci. 2010, 27, 915–922. [CrossRef]
17. Clayton, C.A.; Perritt, R.L.; Pellizzari, E.D.; Thomas, K.W.; Whitmore, R.W.; Ozkaynak, H.; Spengler, J.D.
Particle Total Exposure Assessment Methodology (PTEAM) study: Distributions of aerosol and elemental
concentrations in personal, indoor, and outdoor air samples in a southern California community. J. Expo.
Anal. Environ. Epidemiol. 1993, 3, 227–250. [PubMed]
18. Jantunen, M.J.; Hänninen, O.; Katsouyanni, K.; Knöppel, H.; Kuenzli, N.; Lebret, E.; Maroni, M.; Saarela, K.;
Srám, R.; Zmirou, D. Air pollution exposure in European cities: The “EXPOLIS” study. J. Expo. Anal. Environ.
Epidemiol. 1998, 8, 495–518.
19. Lioy, P.J.; Waldman, J.M.; Buckley, T.; Butler, J.; Pietarinen, C. The personal, indoor and outdoor concentrations
of PM-10 measured in an industrial community during the winter. Atmos. Environ. Part B Urban Atmos. 1990,
24, 57–66. [CrossRef]
20. SciVal. Available online: https://www.scival.com/ (accessed on 7 May 2017).
21. Karagulian, F.; Belis, C.A.; Dora, C.F.C.; Prüss-Ustün, M.A.; Bonjour, S.; Adair-Rohani, H.; Amann, M.
Contributions to cities’ ambient particulate matter (PM): A systematic review of local source contributions at
global level. Atmos. Environ. 2015, 120, 475–483. [CrossRef]
22. Biswas, P.; Wu, C.-Y. Nanoparticles and the Environment. J. Air Waste Manag. Assoc. 2005, 55, 708–746.
[CrossRef] [PubMed]
23. Brauer, M.; Amann, M.; Burnett, R.T.; Cohen, A.; Dentener, F.; Ezzati, M.; Henderson, S.B.; Krzyzanowski, M.;
Martin, R.V.; Dingenen, R.V.; et al. Exposure Assessment for Estimation of the Global Burden of Disease
Attributable to Outdoor Air Pollution. Environ. Sci. Technol. 2012, 46, 652–660. [CrossRef] [PubMed]
24. Murr, L.E.; Garza, K.M. Natural and anthropogenic environmental nanoparticulates: Their microstructural
characterization and respiratory health implications. Atmos. Environ. 2009, 43, 2683–2692. [CrossRef]
25. Wangchuk, T.; He, C.; Dudzinska, M.R.; Morawska, L. Seasonal variations of outdoor air pollution and factors
driving them in the school environment in rural Bhutan. Atmos. Environ. 2015, 113, 151–158. [CrossRef]
26. Monn, C. Exposure assessment of air pollutants: A review on spatial heterogeneity and indoor/outdoor/personal
exposure to suspended particulate matter, nitrogen dioxide and ozone. Atmos. Environ. 2001, 35, 1–32. [CrossRef]
27. Abt, E.; Suh, H.H.; Allen, G.; Koutrakis, P. Characterization of indoor particle sources: A study conducted in
the metropolitan Boston area. Environ. Health Perspect. 2000, 108, 35–44. [CrossRef] [PubMed]
28. Jones, N.C.; Thornton, C.A.; Mark, D.; Harrison, R.M. Indoor/outdoor relationships of particulate matter in
domestic homes with roadside, urban and rural locations. Atmos. Environ. 2000, 34, 2603–2612. [CrossRef]
29. Lai, H.K.; Kendall, M.; Ferrier, H.; Lindup, I.; Alm, S.; Hänninen, O.; Jantunen, M.; Mathys, P.; Colvile, R.;
Ashmore, M.R.; et al. Personal exposures and microenvironment concentrations of PM2.5 , VOC, NO2 and
CO in Oxford, UK. Atmos. Environ. 2004, 38, 6399–6410. [CrossRef]
30. Yocom, J.E. A Critical Review. J. Air Pollut. Control Assoc. 1982, 32, 500–520. [CrossRef]
31. Nazaroff, W.W. Indoor particle dynamics. Indoor Air 2004, 14, 175–183. [CrossRef] [PubMed]
32. Fuoco, F.C.; Stabile, L.; Buonanno, G.; Trassiera, C.V.; Massimo, A.; Russi, A.; Mazaheri, M.; Morawska, L.;
Andrade, A. Indoor Air Quality in Naturally Ventilated Italian Classrooms. Atmosphere 2015, 6, 1652–1675.
[CrossRef]
Atmosphere 2017, 8, 136 16 of 18

33. Rivas, I.; Viana, N.; Morento, T.; Pandolfi, M.; Amato, F.; Reche, C.; Bouso, L.; Alvarez-Pedrerol, M.;
Alastuey, A.; Sunyer, J.; et al. Child exposure to indoor and outdoor air pollutants in schools in Barcelona,
Spain. Environ. Int. 2014, 69, 200–212. [CrossRef] [PubMed]
34. Tofful, L.; Perrino, C. Chemical Composition of Indoor and Outdoor PM2.5 in Three Schools in the City of
Rome. Atmosphere 2015, 6, 1422–1443. [CrossRef]
35. Mainka, A.; Zajusz-Zubek, E.; Kaczmarek, K. PM2.5 in Urban and Rural Nursery Schools in Upper Silesia,
Poland: Trace Elements Analysis. Int. J. Environ. Res. Public Health 2015, 12, 7990–8008. [CrossRef] [PubMed]
36. Lachenmyer, C. Urban Measurements of Outdoor-Indoor PM2.5 Concentrations and Personal Exposure in
the Deep South. Part I. Pilot Study of Mass Concentrations for Nonsmoking Subjects. Aerosol Sci. Technol.
2000, 32, 34–51. [CrossRef]
37. Riesenfeld, E.; Chalupa, D.; Gibb, F.R.; Oberdörster, G.; Gelein, R.; Morrow, P.E.; Utell, M.J.; Frampton, M.W.
Ultrafine Particle Concentrations in a Hospital. Inhal. Toxicol. 2000, 12, 83–94. [CrossRef] [PubMed]
38. Hoek, G.; Kos, G.; Harrison, R.; de Hartog, J.; Meliefste, K.; ten Brink, H.; Katsouyanni, K.; Karakatsani, A.;
Lianou, M.; Kotronarou, A.; et al. Indoor–outdoor relationships of particle number and mass in four
European cities. Atmos. Environ. 2008, 42, 156–169. [CrossRef]
39. Hussein, T.; Hämeri, K.; Aalto, P.P.; Paatero, P.; Kulmala, M. Modal structure and spatial–temporal variations
of urban and suburban aerosols in Helsinki—Finland. Atmos. Environ. 2005, 39, 1655–1668. [CrossRef]
40. Meng, Q.Y.; Turpin, B.J.; Korn, L.; Weisel, C.P.; Morandi, M.; Colome, S.; Zhang, J.; Stock, T.; Spektor, D.;
Winer, A. Influence of ambient (outdoor) sources on residential indoor and personal PM2.5 concentrations:
Analyses of RIOPA data. J. Expo. Sci. Environ. Epidemiol. 2005, 15, 17–28. [CrossRef] [PubMed]
41. Brunekreef, B.; Janssen, N.A.H.; de Hartog, J.J.; Oldenwening, M.; Meliefste, K.; Hoek, G.; Lanki, T.;
Timonen, K.L.; Vallius, M.; Pekkanen, J.; et al. Personal, indoor, and outdoor exposures to PM2.5 and its
components for groups of cardiovascular patients in Amsterdam and Helsinki. Res. Rep. Health Eff. Inst.
2005, 127, 1–70.
42. Shilton, V.; Giess, P.; Mitchell, D.; Williams, C. The Relationships between Indoor and Outdoor Respirable
Particulate Matter: Meteorology, Chemistry and Personal Exposure. Indoor Built Environ. 2002, 11, 266–274.
[CrossRef]
43. Zhou, B.; Zhao, B.; Guo, X.; Chen, R.; Kan, H. Investigating the geographical heterogeneity in PM10 -mortality
associations in the China Air Pollution and Health Effects Study (CAPES): A potential role of indoor exposure
to PM10 of outdoor origin. Atmos. Environ. 2013, 75, 217–223. [CrossRef]
44. Zhao, B.; Wu, J. Particle deposition in indoor environments: Analysis of influencing factors. J. Hazard. Mater.
2007, 147, 439–448. [CrossRef] [PubMed]
45. Nazir, R.; Shaheen, N.; Shah, M.H. Indoor/outdoor relationship of trace metals in the atmospheric particulate
matter of an industrial area. Atmos. Res. 2011, 101, 765–772. [CrossRef]
46. Krugly, E.; Martuzevicius, D.; Sidaraviciute, R.; Ciuzas, D.; Prasauskas, T.; Kauneliene, V.; Stasiulaitiene, I.;
Kliucininkas, L. Characterization of particulate and vapor phase polycyclic aromatic hydrocarbons in indoor
and outdoor air of primary schools. Atmos. Environ. 2014, 82, 298–306. [CrossRef]
47. Halsall, C.J.; Maher, B.A.; Karloukovski, V.V.; Shah, P.; Watkins, S.J. A novel approach to investigating
indoor/outdoor pollution links: Combined magnetic and PAH measurements. Atmos. Environ. 2008,
42, 8902–8909. [CrossRef]
48. Alves, C.; Duarte, M.; Ferreira, M.; Alves, A.; Almeida, A.; Cunha, Â. Air quality in a school with dampness
and mould problems. Air Qual. Atmos. Health 2016, 9, 107–115. [CrossRef]
49. Pagel, É.C.; Reis, N.C.; Alvarez, C.E.; Santos, J.M.; Conti, M.M.; Boldrini, R.S.; Kerr, A.S. Characterization of
the indoor particles and their sources in an Antarctic research station. Environ. Monit. Assess. 2016, 188, 167.
[CrossRef] [PubMed]
50. Saraga, D.; Pateraki, S.; Papadopoulos, A.; Vasilakos, C.; Maggos, T. Studying the indoor air quality in three
non-residential environments of different use: A museum, a printery industry and an office. Build. Environ.
2011, 46, 2333–2341. [CrossRef]
51. Polednik, B. Particulate matter and student exposure in school classrooms in Lublin, Poland. Environ. Res.
2013, 120, 134–139. [CrossRef] [PubMed]
52. Worobiec, A.; Samek, L.; Krata, A.; van Meel, K.; Krupinska, B.; Stefaniak, E.A.; Karaszkiewicz, P.;
van Grieken, R. Transport and deposition of airborne pollutants in exhibition areas located in historical
buildings–study in Wawel Castle Museum in Cracow, Poland. J. Cult. Herit. 2010, 11, 354–359. [CrossRef]
Atmosphere 2017, 8, 136 17 of 18

53. Chatoutsidou, S.E.; Ondráček, J.; Tesar, O.; Tørseth, K.; Ždímal, V.; Lazaridis, M. Indoor/outdoor particulate
matter number and mass concentration in modern offices. Build. Environ. 2015, 92, 462–474. [CrossRef]
54. Diapouli, E.; Eleftheriadis, K.; Karanasiou, A.A.; Vratolis, S.; Hermansen, O.; Colbeck, I.; Lazaridis, M. Indoor
and Outdoor Particle Number and Mass Concentrations in Athens. Sources, Sinks and Variability of Aerosol
Parameters. Aerosol Air Qual. Res. 2011, 11, 632–642. [CrossRef]
55. Vicente, E.D.; Ribeiro, J.P.; Custódio, D.; Alves, C.A. Assessment of the indoor air quality in copy centres at
Aveiro, Portugal. Air Qual. Atmos. Health 2017, 10, 117–127. [CrossRef]
56. Braniš, M.; Řezáčová, P.; Domasová, M. The effect of outdoor air and indoor human activity on mass
concentrations of PM10 , PM2.5 , and PM1 in a classroom. Environ. Res. 2005, 99, 143–149. [CrossRef]
[PubMed]
57. Goyal, R.; Kumar, P. Indoor-outdoor concentrations of particulate matter in nine microenvironments of
a mix-use commercial building in megacity Delhi. Air Qual. Atmos. Health 2013, 6, 747–757. [CrossRef]
58. Liu, Y.; Chen, R.; Shen, X.; Mao, X. Wintertime indoor air levels of PM10 , PM2.5 and PM1 at public places and
their contributions to TSP. Environ. Int. 2004, 30, 189–197. [CrossRef]
59. Schembari, A.; Triguero-Mas, M.; de Nazelle, A.; Davdand, P.; Vrijheid, M.; Cirach, M.; Martinez, D.;
Figueras, F.; Querol, X.; Basagaña, X. Personal, indoor and outdoor air pollution levels among pregnant
women. Atmos. Environ. 2013, 64, 287–295. [CrossRef]
60. Wheeler, A.J.; Wallace, L.A.; Kearney, J.; van Ryswyk, K.; You, H.; Kulka, R.; Brook, J.R.; Xu, X. Personal,
Indoor, and Outdoor Concentrations of Fine and Ultrafine Particles Using Continuous Monitors in Multiple
Residences. Aerosol Sci. Technol. 2011, 45, 1078–1089. [CrossRef]
61. Sajani, S.Z.; Ricciardelli, I.; Trentini, A.; Bacco, D.; Maccone, C.; Castellazzi, S.; Lauriola, P.; Poluzzi, V.;
Harrison, R.M. Is particulate air pollution at the front door a good proxy of residential exposure?
Environ. Pollut. 2016, 213, 347–358. [CrossRef] [PubMed]
62. Weichenthal, S.; Dufresne, A.; Infante-Rivard, C.; Joseph, L. Indoor ultrafine particle exposures and home
heating systems: A cross-sectional survey of Canadian homes during the winter months. J. Expo. Sci.
Environ. Epidemiol. 2006, 17, 288–297. [CrossRef] [PubMed]
63. Colbeck, I.; Nasir, Z.A.; Ali, Z. Characteristics of indoor/outdoor particulate pollution in urban and rural
residential environment of Pakistan. Indoor Air 2010, 20, 40–51. [CrossRef] [PubMed]
64. Majewski, G.; Kociszewska, K.; Rogula-Kozłowska, W.; Pyta, H.; Rogula-Kopiec, P.; Mucha, W.; Pastuszka, J.
Submicron Particle-Bound Mercury in University Teaching Rooms: A Summer Study from Two Polish Cities.
Atmosphere 2016, 7, 117. [CrossRef]
65. Barthel, M.; Pedan, V.; Hahn, O.; Rothhardt, M.; Bresch, H.; Jann, O.; Seeger, S. XRF-Analysis of Fine and
Ultrafine Particles Emitted from Laser Printing Devices. Environ. Sci. Technol. 2011, 45, 7819–7825. [CrossRef]
[PubMed]
66. Chatoutsidou, S.E.; Serfozo, N.; Glytsos, T.; Lazaridis, M. Multi-zone measurement of particle concentrations
in a HVAC building with massive printer emissions: Influence of human occupation and particle transport
indoors. Air Qual. Atmos. Health 2017, 1–15. [CrossRef]
67. Lee, C.-W.; Hsu, D.-J. Measurements of fine and ultrafine particles formation in photocopy centers in Taiwan.
Atmos. Environ. 2007, 41, 6598–6609. [CrossRef]
68. Diapouli, E.; Chaloulakou, A.; Mihalopoulos, N.; Spyrellis, N. Indoor and outdoor PM mass and number
concentrations at schools in the Athens area. Environ. Monit. Assess. 2008, 136, 13–20. [CrossRef] [PubMed]
69. Kuhlbusch, T.A.; Asbach, C.; Fissan, H.; Göhler, D.; Stintz, M. Nanoparticle exposure at nanotechnology
workplaces: A review. Part. Fibre Toxicol. 2011, 8, 22. [CrossRef] [PubMed]
70. Dahm, M.M.; Evans, D.E.; Schubauer-Berigan, M.K.; Birch, M.E.; Deddens, J.A. Occupational Exposure
Assessment in Carbon Nanotube and Nanofiber Primary and Secondary Manufacturers: Mobile
Direct-Reading Sampling. Ann. Occup. Hyg. 2013, 57, 328–344. [PubMed]
71. Xu, H.; Guinot, B.; Shen, Z.; Ho, K.F.; Niu, X.; Xiao, S.; Huang, R.J.; Cao, J. Characteristics of Organic and
Elemental Carbon in PM2.5 and PM0.25 in Indoor and Outdoor Environments of a Middle School: Secondary
Formation of Organic Carbon and Sources Identification. Atmosphere 2015, 6, 361–379. [CrossRef]
72. Amaral, S.S.; De Carvalho, J.A.; Costa, M.A.M.; Pinheiro, C. An Overview of Particulate Matter Measurement
Instruments. Atmosphere 2015, 6, 1327–1345. [CrossRef]
Atmosphere 2017, 8, 136 18 of 18

73. Chow, J.C.; Doraiswamy, P.; Watson, J.G.; Chen, L.-W.A.; Ho, S.S.H.; Sodeman, D.A. Advances in Integrated
and Continuous Measurements for Particle Mass and Chemical Composition. J. Air Waste Manag. Assoc.
2008, 58, 141–163. [CrossRef] [PubMed]
74. Marple, V.A.; Rubow, K.L.; Turner, W.; Spengler, J.D. Low Flow Rate Sharp Cut Impactors for Indoor Air
Sampling: Design and Calibration. JAPCA 1987, 37, 1303–1307. [CrossRef] [PubMed]
75. Crist, K.C.; Liu, B.; Kim, M.; Deshpande, S.R.; John, K. Characterization of fine particulate matter in Ohio:
Indoor, outdoor, and personal exposures. Environ. Res. 2008, 106, 62–71. [CrossRef] [PubMed]
76. Kuo, H.-W.; Shen, H.-Y. Indoor and outdoor PM2.5 and PM10 concentrations in the air during a dust storm.
Build. Environ. 2010, 45, 610–614. [CrossRef]
77. Steinle, S.; Reis, S.; Sabel, C.E.; Semple, S.; Twigg, M.M.; Braban, C.F.; Leeson, S.R.; Heal, M.R.;
Harrison, D.; Lin, C.; et al. Personal exposure monitoring of PM2.5 in indoor and outdoor microenvironments.
Sci. Total Environ. 2015, 508, 383–394. [CrossRef] [PubMed]
78. United States Environmental Protection Agency (US EPA). 2012 National Ambient Air Quality Standards
(NAAQS) for Particulate Matter (PM). Available online: https://www.epa.gov/pm-pollution/2012-national-
ambient-air-quality-standards-naaqs-particulate-matter-pm (assessed on 16 May 2017).
79. Challoner, A.; Gill, L. Indoor/outdoor air pollution relationships in ten commercial buildings: PM2.5 and
NO2 . Build. Environ. 2014, 80, 159–173. [CrossRef]
80. Weisel, C.P.; Zhang, J.; Turpin, B.J.; Morandi, M.T.; Colome, S.; Stock, T.H.; Spektor, D.M.; Korn, L.; Winer, A.;
Alimokhtari, S. Relationship of Indoor, Outdoor and Personal Air (RIOPA) study: Study design, methods and
quality assurance/control results. J. Expo. Sci. Environ. Epidemiol. 2005, 15, 123–137. [CrossRef] [PubMed]
81. Koistinen, K.J.; Kousa, A.; Tenhola, V.; Hänninen, O.; Jantunen, M.J.; Oglesby, L.; Kuenzli, N.; Georgoulis, L.
Fine Particle (PM25 ) Measurement Methodology, Quality Assurance Procedures, and Pilot Results of the
EXPOLIS Study. J. Air Waste Manag. Assoc. 1999, 49, 1212–1220. [CrossRef] [PubMed]
82. Guo, H.; Morawska, L.; He, C.; Gilbert, D. Impact of ventilation scenario on air exchange rates and on
indoor particle number concentrations in an air-conditioned classroom. Atmos. Environ. 2008, 42, 757–768.
[CrossRef]
83. Fromme, H.; Diemer, J.; Dietrich, S.; Cyrys, J.; Heinrich, J.; Lang, W.; Kiranoglu, M.; Twardella, D. Chemical
and morphological properties of particulate matter (PM10 , PM2.5 ) in school classrooms and outdoor air.
Atmos. Environ. 2008, 42, 6597–6605. [CrossRef]
84. Diapouli, E.; Chaloulakou, A.; Koutrakis, P. Estimating the concentration of indoor particles of outdoor
origin: A review. J. Air Waste Manag. Assoc. 2013, 63, 1113–1129. [CrossRef] [PubMed]
85. Kousa, A.; Oglesby, L.; Koistinen, K.; Künzli, N.; Jantunen, M. Exposure chain of urban air
PM2.5 —Associations between ambient fixed site, residential outdoor, indoor, workplace and personal
exposures in four European cities in the EXPOLIS-study. Atmos. Environ. 2002, 36, 3031–3039. [CrossRef]
86. Snyder, E.G.; Watkins, T.H.; Solomon, P.A.; Thomas, E.D.; Williams, R.W.; Hagler, G.S.W.; Shelow, D.;
Hindin, D.A.; Kilaru, V.J.; Preuss, P.W. The Changing Paradigm of Air Pollution Monitoring.
Environ. Sci. Technol. 2013, 47, 11369–11377. [CrossRef] [PubMed]
87. Oliveira, M.; Slezakova, K.; Madureira, J.; de Oliveira Fernandes, E.; Delerue-Matos, C.; Morais, S.;
do Carmo Pereira, M. Polycyclic aromatic hydrocarbons in primary school environments: Levels and
potential risks. Sci. Total Environ. 2017, 575, 1156–1167. [CrossRef] [PubMed]
88. Romagnoli, P.; Balducci, C.; Perilli, M.; Vichi, F.; Imperiali, A.; Cecinato, A. Indoor air quality at life and work
environments in Rome, Italy. Environ. Sci. Pollut. Res. 2016, 23, 3503–3516. [CrossRef] [PubMed]
89. Błaszczyk, E.; Rogula-Kozłowska, W.; Klejnowski, K.; Fulara, I.; Mielżyńska-Švach, D. Polycyclic aromatic
hydrocarbons bound to outdoor and indoor airborne particles (PM2.5 ) and their mutagenicity and
carcinogenicity in Silesian kindergartens, Poland. Air Qual. Atmos. Health 2017, 10, 389–400. [CrossRef]
[PubMed]
90. Rogula-Kopiec, P.; Rogula-Kozłowska, W.; Kozielska, B.; Sówka, I. PAH Concentrations Inside a Wood
Processing Plant and the Indoor Effects of Outdoor Industrial Emissions. Pol. J. Environ. Stud. 2015, 24, 11–17.
[CrossRef]

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

Das könnte Ihnen auch gefallen